Fact-checked by Grok 2 weeks ago

Linear response function

The linear response function, also known as the response function or , is a in and quantum physics that describes the linear change in the expectation value of an in a subjected to a weak, time-dependent external . It quantifies how the system deviates from , capturing relationships between the perturbation and the induced response while ensuring principles like causality and are satisfied through its retarded nature. This framework assumes the perturbation is small enough that higher-order effects are negligible, allowing the response to be proportional to the perturbation's . Developed primarily in the mid-20th century, linear response theory was formalized by Ryogo Kubo in his seminal 1957 paper, which provided a general statistical-mechanical basis for calculating irreversible processes and transport coefficients from equilibrium correlation functions. Kubo's approach bridges microscopic dynamics to macroscopic phenomena, showing that the response is directly tied to fluctuations in the unperturbed system via the . In , the theory applies to systems at finite temperatures, using the to average over equilibrium states. The cornerstone of the theory is the Kubo formula, which expresses the response function \chi_{AB}(t - t') between observables A and B as \chi_{AB}(t - t') = -i \theta(t - t') \langle [A(t), B(t')] \rangle_0, where \theta is the enforcing , and the brackets denote the equilibrium average of the in the . In frequency space, this becomes a function whose real part relates to reactive effects and imaginary part to dissipation, enabling computations via transforms of correlation functions. For classical systems, an analogous replaces the . Applications of linear response functions span , including the calculation of electrical \sigma(\omega) = \frac{e^2}{i\omega} \chi_{jj}(\omega) from current-current correlations, responses in materials, and magnetic susceptibilities. It is also essential for interpreting experiments like optical absorption, neutron scattering, and angle-resolved photoemission spectroscopy (ARPES), where external probes reveal intrinsic material properties. Extensions to nonequilibrium and open quantum systems further broaden its utility in modern fields like and nanoscale transport.

Basic Concepts

Definition

Linear response theory provides a fundamental framework in and physics for predicting the behavior of physical systems subjected to weak external , such as electric or magnetic fields. It assumes that the system's response, measured through changes in values of observables, is linearly proportional to the strength of the , which is valid for small disturbances that do not significantly alter the state. The core concept is encapsulated in the response function \chi_{AB}(t - t'), which quantifies the change in the expectation value of A at time t, denoted \delta \langle A(t) \rangle, due to a \delta h_B(t') in the conjugate field h_B at an earlier or simultaneous time t'. This relationship is expressed as \delta \langle A(t) \rangle = \int_{-\infty}^{\infty} \chi_{AB}(t - t') \, \delta h_B(t') \, dt', where the integral accounts for the superposition of responses from all perturbation times, assuming time-translation invariance in the unperturbed system. A key property of the response function is causality, which dictates that \chi_{AB}(t - t') = 0 for t < t', ensuring that the system's response cannot precede the applied and thus preserving the directional flow of time in physical processes. This theoretical framework originated in the 1950s within the development of nonequilibrium statistical mechanics, with pivotal contributions from and collaborators, who formalized the connection between linear responses and equilibrium fluctuations.

Linear Approximation

The linear approximation in response theory assumes that the perturbation to the system's Hamiltonian, typically expressed as \delta H = -h(t) B where h(t) is a small external field and B is an operator coupling to it, is much weaker than the unperturbed Hamiltonian H_0, i.e., \delta H \ll H_0. This allows the system's response to be described by retaining only the first-order term in the perturbation expansion, neglecting higher-order contributions that become negligible under this condition. The change in the expectation value of an observable A, denoted \delta \langle A \rangle, can be expanded around the equilibrium state using a Taylor series in the perturbation strength h: \delta \langle A \rangle \approx \left. \frac{\partial \langle A \rangle}{\partial h} \right|_{h=0} \delta h, which captures the first-order linear response proportional to \delta h. This approximation holds under conditions of weak coupling between the perturbation and the system, where the field h(t) induces only small deviations from equilibrium, and for perturbation durations short compared to the system's relaxation times, ensuring the response remains causal and does not allow full re-equilibration. It applies equally to both classical and quantum systems, as the underlying perturbative expansion is general. When perturbations are strong such that \delta H \sim H_0, the linear approximation breaks down, necessitating nonlinear response theory to account for higher-order effects.

Mathematical Formalism

Time-Domain Formulation

In the time domain, the linear response of an observable A(t) to a small perturbation coupled to another observable B via an external field h_B(t) is expressed as the change in the equilibrium average: \delta \langle A(t) \rangle = \int_{-\infty}^t \chi_{AB}(t - t') \, \delta h_B(t') \, dt', where \chi_{AB}(\tau) is the response function, defined for \tau > 0 as the functional derivative \chi_{AB}(\tau) = \left. \frac{\delta \langle A(t + \tau) \rangle}{\delta h_B(t)} \right|_{\rm eq}, evaluated at equilibrium with \delta h_B = 0. This formulation assumes the system starts in equilibrium before the perturbation at t' = -\infty, and the upper limit of integration enforces causality, as the response at time t depends only on past and present perturbations. In classical , the response function \chi_{AB}(\tau) is directly related to time-correlation functions of fluctuations. Specifically, \chi_{AB}(\tau) = \beta \langle \dot{B}(0) A(\tau) \rangle, where \beta = 1/(k_B T), k_B is Boltzmann's constant, T is , \dot{B} denotes the time derivative of B, and the angular brackets indicate an ensemble average. This relation, derived from the of the phase-space , links the nonequilibrium response to measurable fluctuations, providing a practical way to compute \chi_{AB} from simulations or experiments. The response function \chi_{AB}(\tau) is a retarded function, vanishing for \tau < 0 due to causality, and in stationary systems—where the Hamiltonian is time-independent—it exhibits time-translation invariance, depending only on the time difference \tau = t - t'. These properties ensure that the response is physically meaningful, with \chi_{AB}(\tau) decaying to zero as \tau \to \infty in dissipative systems, reflecting the system's return to equilibrium. Regarding units and dimensions, \chi_{AB}(\tau) has the dimensions of the observable A divided by the field strength h_B, mirroring the units of physical susceptibilities such as electric permittivity (relating polarization to electric field) or magnetic permeability (relating magnetization to magnetic field). For instance, in electrostatics, the time-domain response function for dielectric susceptibility connects to the material's polarizability under time-varying fields.

Frequency-Domain Formulation

The frequency-domain formulation of the linear response function is obtained by taking the Fourier transform of the time-domain response function \chi_{AB}(\tau), which relates the change in the expectation value of observable A to a perturbation coupled to observable B. Specifically, the frequency-dependent response function is defined as \chi_{AB}(\omega) = \int_0^\infty \chi_{AB}(\tau) \, e^{i \omega \tau} \, d\tau, where \tau = t - t' is the time lag, and the Fourier transform convention ensures compatibility with causal responses. In this representation, the induced response in frequency space is given by \delta \langle A(\omega) \rangle = \chi_{AB}(\omega) \, \delta h_B(\omega), where \delta h_B(\omega) is the Fourier transform of the perturbing field. This transformation simplifies the analysis of systems driven by harmonic perturbations, converting convolutions in time into algebraic multiplications in frequency. The frequency-domain response function \chi_{AB}(\omega) exhibits key properties arising from the underlying causality of the system, which was established in the time-domain formulation. Causality implies that \chi_{AB}(\omega) is analytic in the upper half of the complex frequency plane (Im \omega > 0), ensuring no unphysical anticipatory responses. Additionally, the real part \operatorname{Re}[\chi_{AB}(\omega)] is an even function of \omega, while the imaginary part \operatorname{Im}[\chi_{AB}(\omega)] is odd, reflecting the symmetry of the underlying equilibrium correlations. These properties lead to the Kramers-Kronig relations, which interconnect the real and imaginary parts through Hilbert transforms: \operatorname{Re}[\chi_{AB}(\omega)] = \frac{2}{\pi} \mathcal{P} \int_0^\infty \frac{\omega' \operatorname{Im}[\chi_{AB}(\omega')]}{\omega'^2 - \omega^2} \, d\omega', \operatorname{Im}[\chi_{AB}(\omega)] = -\frac{2\omega}{\pi} \mathcal{P} \int_0^\infty \frac{\operatorname{Re}[\chi_{AB}(\omega')]}{\omega'^2 - \omega^2} \, d\omega', where \mathcal{P} denotes the Cauchy principal value. These relations, derived solely from causality and stability, allow the computation of one part from the other, often using experimentally measured absorption data to infer dispersion. For (AC) fields, the frequency-domain formulation is particularly useful in describing steady-state behaviors under harmonic driving. Consider a sinusoidal \delta h_B(t) = \operatorname{Re}[\delta h_B(\omega) e^{-i \omega t}]; the resulting response is \delta \langle A(t) \rangle = \operatorname{Re}[\chi_{AB}(\omega) \delta h_B(\omega) e^{-i \omega t}], which includes a shift relative to the driving . In this , the real part \operatorname{Re}[\chi_{AB}(\omega)] governs the in-phase component of the response, corresponding to dissipative processes such as energy absorption, while the imaginary part \operatorname{Im}[\chi_{AB}(\omega)] governs the out-of-phase (quadrature) component, associated with reactive or dispersive effects like without net loss. The rate is proportional to \operatorname{Re}[\chi_{AB}(\omega)] |\delta h_B(\omega)|^2, highlighting the role of the in-phase response in thermodynamic consistency. In the static limit as \omega \to 0, the frequency-domain response function reduces to the susceptibility \chi_{AB}(0), which quantifies the isothermal response of the system to a constant field, such as the in a static . This limit connects the dynamic formulation back to static thermodynamic derivatives, like \chi_{AB}(0) = \left( \frac{\partial \langle A \rangle}{\partial h_B} \right)_{T,V}, and is real and positive for stable systems.

Quantum Formulation

Kubo Formula

The Kubo formula expresses the linear response function in as a of operators in , providing a foundational tool for calculating transport coefficients and susceptibilities in many-body systems. Introduced by Ryogo Kubo in 1957, it establishes linear response theory as a pillar of quantum many-body physics by linking perturbations to equilibrium fluctuations via the of the unperturbed . In the time domain, the response function \chi_{AB}(t - t') relating the expectation value of observable A to a perturbation coupled to observable B takes the form of a retarded commutator, with thermal equilibrium averages denoted by \langle \cdot \rangle_0: \chi_{AB}(t - t') = -\frac{i}{\hbar} \theta(t - t') \langle [A(t), B(t')] \rangle_0, where \theta is the Heaviside step function ensuring causality, [ \cdot, \cdot ] is the quantum commutator, and the Heisenberg-picture operators evolve under the unperturbed Hamiltonian. This expression captures the causal propagation of disturbances in quantum systems, distinguishing it from classical response functions by incorporating non-commutativity. Thermal corrections arise implicitly through the equilibrium average, but the primary structure relies on this commutator form. In the , the of the retarded response yields \chi_{AB}(\omega) = -\frac{i}{\hbar} \int_0^\infty dt \, e^{i \omega t} \langle [A(t), B(0)] \rangle_0, which exhibits the analytic properties required for , such as satisfaction of the Kramers-Kronig relations. For the dissipative component, which governs energy absorption and is the imaginary part \operatorname{Im} \chi_{AB}(\omega) in standard conventions, the Kubo formula connects to symmetrized correlations via the : \operatorname{Im} \chi_{AB}(\omega) = \frac{1 - e^{-\beta \hbar \omega}}{2 \hbar} \int_{-\infty}^\infty dt \, e^{i \omega t} \langle \{A(t), B(0)\}/2 \rangle_0, where \beta = 1/(k_B T) is the inverse temperature. This form highlights the role of quantum fluctuations in dissipation, with the factor (1 - e^{-\beta \hbar \omega}) encoding Bose-Einstein statistics for bosonic observables. Equivalently, using the commutator, the dissipative part can be expressed in terms of unsymmetrized correlations, but the symmetrized form emphasizes the direct link to equilibrium noise spectra. The distinction between the retarded (commutator-based) and canonical (symmetrized) response functions is central in the quantum context: the retarded form enforces and is directly tied to the Kubo formula's perturbative origin, while the symmetrized form relates to measurable fluctuations and is obtained via the . This duality arises from the Kubo-Martin-Schwinger () condition, which characterizes states by imposing periodicity in on correlation functions: for operators A and B, \langle A(t) B(0) \rangle_0 = \langle B(0) A(t + i \beta \hbar) \rangle_0. The KMS condition ensures the equivalence between commutator and symmetrized expressions, enabling the computation of response functions from thermal averages without explicit perturbation theory. It underpins the thermal factors in the dissipative form and extends the Kubo formula's applicability to finite temperatures.

Derivation of Kubo Formula

The derivation of the Kubo formula begins with the quantum mechanical description of a system in thermal equilibrium perturbed by an external field, governed by the von Neumann equation for the density operator \rho(t): i \hbar \frac{d \rho(t)}{dt} = [H_0 + \delta H(t), \rho(t)], where H_0 is the unperturbed Hamiltonian, and the perturbation is \delta H(t) = - h(t) B with h(t) the external field and B the coupling operator. In the linear response regime, the density operator is expanded perturbatively as \rho(t) = \rho_{\rm eq} + \rho^{(1)}(t) + \cdots, where \rho_{\rm eq} is the equilibrium density matrix satisfying [H_0, \rho_{\rm eq}] = 0. Substituting into the von Neumann equation and retaining terms to first order in the perturbation yields i \hbar \frac{d \rho^{(1)}(t)}{dt} = [H_0, \rho^{(1)}(t)] + [\delta H(t), \rho_{\rm eq}], with the solution, assuming the system starts in equilibrium at t \to -\infty, \rho^{(1)}(t) = -\frac{i}{\hbar} \int_{-\infty}^t dt' \, e^{-i H_0 (t - t') / \hbar} [\delta H(t'), \rho_{\rm eq}] e^{i H_0 (t - t') / \hbar}. This integral form accounts for the evolution under H_0 between the perturbation at t' and observation at t. The linear change in the expectation value of an observable A is then \delta \langle A(t) \rangle = {\rm Tr} [A \rho^{(1)}(t)]. Substituting the expression for \rho^{(1)}(t) and \delta H(t') = - h(t') B gives \delta \langle A(t) \rangle = -\frac{i}{\hbar} \int_{-\infty}^t dt' \, h(t') \, {\rm Tr} \left\{ \rho_{\rm eq} \left[ A, e^{-i H_0 (t - t') / \hbar} B e^{i H_0 (t - t') / \hbar} \right] \right\}. This expresses the response in terms of the average of a . To obtain the causal response function, the is employed, where operators evolve as A_I(t) = e^{i H_0 t / \hbar} A e^{-i H_0 t / \hbar} and similarly for B_I(t). The trace becomes time-translation invariant, leading to the linear response function \chi_{AB}(t - t') = -\frac{i}{\hbar} \theta(t - t') \, {\rm Tr} \left\{ \rho_{\rm eq} [A_I(t), B_I(t')] \right\}, with \theta the ensuring causality. Thus, \delta \langle A(t) \rangle = \int_{-\infty}^\infty dt' \, \chi_{AB}(t - t') h(t'). This derivation assumes weak where higher-order terms are negligible, and the external field varies slowly compared to microscopic timescales, without invoking a Markovian . The state \rho_{\rm eq} is preserved in the absence of the perturbation.

Extensions and Applications

Nonequilibrium Response

In nonequilibrium systems, linear response theory extends to perturbations around nonequilibrium steady states (NESS), where external drives maintain persistent currents or fluxes, leading to stationary distributions distinct from . The response functions are defined using correlations evaluated in this NESS, often via the stationary density operator solving the Lindblad or for open . For example, the frequency-dependent \sigma(\omega) in electrically driven systems arises from current-current correlations in the NESS, capturing dissipative transport under weak fields. A Kubo-like formula for NESS generalizes the equilibrium expression by introducing a steady-state Kubo transformation to address operator noncommutativity and the absence of . In stationary nonequilibrium conditions, the response function \chi_{AB}(t,t') retains time-translation invariance, depending only on the difference t - t', analogous to but with NESS averages. However, in transient nonequilibrium scenarios—such as systems evolving after a quench or under time-dependent drives—time-translation invariance breaks, making \chi_{AB}(t,t') explicitly dependent on both observation time t and perturbation time t', which encodes the system's aging or relaxation . For transport applications, linear response yields relations like the current density \mathbf{J} = \sigma \mathbf{E}, applicable near but extendable to moderate nonequilibrium via hybrid methods that integrate Kubo correlation functions with the . These approaches linearize the around a driven , incorporating relaxation-time approximations for , to compute coefficients such as electrical or in inhomogeneous or weakly coupled systems. Recent advances as of 2024 have applied these frameworks to systems and non-Hermitian quantum circuits, enabling analysis of collective behaviors and non-unitary dynamics in driven environments. Limitations arise in far-from-equilibrium regimes, where strong drives like high induce nonlinearities, such as heating or of the relaxation-time , invalidating the linear expansion and necessitating nonlinear or full kinetic theories.

Fluctuation-Dissipation Theorem

The fluctuation-dissipation theorem (FDT) establishes a fundamental connection between the linear response of a physical system to external perturbations and the spontaneous fluctuations inherent in its equilibrium state, providing a deep link between dissipation and noise in both classical and quantum regimes. In the context of linear response theory, the theorem demonstrates that the dissipative part of the response function, which quantifies energy absorption from an applied field, is directly determined by the power spectral density of equilibrium fluctuations of the relevant observable. This relation underscores the principle that any dissipative process must be accompanied by corresponding fluctuations, ensuring consistency with thermodynamic equilibrium. In the classical limit, applicable at high temperatures where \hbar \omega \ll k_B T, the FDT relates the imaginary part of the frequency-domain response function \chi''(\omega) to the power spectral density S_{AA}(\omega) of the equilibrium autocorrelation function \langle A(t) A(0) \rangle as \chi''(\omega) = \frac{\beta \omega}{2} S_{AA}(\omega), where \beta = 1/(k_B T) is the inverse temperature, k_B is Boltzmann's constant, T is the temperature, and \omega is the angular frequency. This form highlights how the strength of dissipation at frequency \omega is proportional to the intensity of fluctuations at the same frequency, scaled by thermal factors. For the quantum case, which accounts for zero-point and thermal excitations, the relation generalizes to \chi''(\omega) = \frac{1 - e^{-\beta \hbar \omega}}{2 \hbar} S_{AA}(\omega), where \hbar is the reduced Planck's constant; here, S_{AA}(\omega) is the Fourier transform of the symmetrized quantum correlation function. This quantum expression reduces to the classical one in the high-temperature limit, as $1 - e^{-\beta \hbar \omega} \approx \beta \hbar \omega, and incorporates quantum statistical effects through the Bose-Einstein factor. The implications of the FDT are profound: it reveals that the linear response of a in is entirely encoded in its intrinsic fluctuations, allowing dissipative properties to be inferred from measurable spectra without external driving. In the static limit (\omega \to 0), the theorem yields the Einstein relation \chi(0) = \beta \langle (\Delta A)^2 \rangle, linking static to equilibrium variance and embodying the fluctuation-response paradigm central to . Historically, the FDT was first derived by Callen and Welton in using a quantum electrodynamic approach to generalized in linear dissipative systems, predating Kubo's formal linear response and unifying thermodynamic fluctuations with dynamic response. This theorem has since become a for interpreting transport coefficients and in condensed systems.

Examples

Harmonic Oscillator

The classical damped driven harmonic oscillator provides a foundational example for illustrating linear response theory, as its equation of motion can be solved exactly to reveal the susceptibility without approximations beyond the linear regime. Consider a particle of mass m subject to a restoring force -kx, a frictional damping term -\gamma \dot{x}, and an external driving force f(t), governed by the equation m \ddot{x} + \gamma \dot{x} + k x = f(t). Here, \omega_0 = \sqrt{k/m} denotes the natural frequency. In the frequency domain, assuming a harmonic drive f(t) = \Re[f(\omega) e^{-i\omega t}] and response x(t) = \Re[x(\omega) e^{-i\omega t}], the susceptibility \chi(\omega) = x(\omega)/f(\omega) is given by \chi(\omega) = \frac{1}{m(\omega_0^2 - \omega^2 - i (\gamma / m) \omega)}, where the imaginary part arises from the damping and encodes dissipative effects. This form is obtained by substituting the Fourier ansatz into the equation of motion and solving for x(\omega), yielding a resonant response peaked near \omega \approx \omega_0, shifted slightly by damping for the absorption (imaginary part of \chi(\omega)) and dispersion (real part). The real part \Re[\chi(\omega)] describes the in-phase (reactive) displacement, while the imaginary part \Im[\chi(\omega)] quantifies out-of-phase motion linked to energy dissipation, proportional to the damping coefficient \gamma; for small \gamma, \Im[\chi(\omega)] exhibits a Lorentzian peak at \omega_0 with width \gamma/m, illustrating how linear response captures resonance phenomena analytically. This mechanical model highlights the pedagogical value of linear response: the damping \gamma directly connects to dissipation via the imaginary susceptibility, enabling prediction of absorption spectra without numerical simulation of the full dynamics. In the quantum case, the harmonic oscillator's position response to an applied force follows the same susceptibility form, derived via the Kubo formula, which expresses the response function as a retarded commutator of position operators in the unperturbed thermal ensemble. For the quantum damped oscillator (modeled via coupling to a bath), the calculation proceeds by evaluating the position-position correlation function, yielding \chi(\omega) identical to the classical result in the linear regime, with quantum corrections manifesting as zero-point fluctuations related to the fluctuation-dissipation theorem. This equivalence underscores the oscillator's utility in bridging classical and quantum linear response, where the same \Re[\chi(\omega)] and \Im[\chi(\omega)] plots reveal dispersion and absorption, now incorporating thermal and quantum noise.

Electrical Conductivity

In linear response theory, electrical quantifies the transport of in response to an applied , serving as a fundamental transport . The \mathbf{J} in a relates linearly to the \mathbf{E} via \mathbf{J}(\omega) = \boldsymbol{\sigma}(\omega) \mathbf{E}(\omega), where \boldsymbol{\sigma}(\omega) is the frequency-dependent tensor. This relation arises from the general framework of linear response, where the perturbation Hamiltonian is H' = -\int d\mathbf{r} \, \mathbf{J}(\mathbf{r}) \cdot \mathbf{A}(\mathbf{r}), with \mathbf{A} the and \mathbf{E} = -\partial \mathbf{A}/\partial t for transverse fields in the gauge. The tensor components \sigma_{\alpha\beta}(\omega) are expressed through the retarded current-current , capturing dissipative and reactive responses in equilibrium systems. The Kubo formula provides the explicit expression for the , derived from the of operators in the . For a of volume V, the tensor elements are given by \sigma_{\alpha\beta}(\omega) = \frac{ie^2}{\omega V} \sum_{\mathbf{k}} \frac{\partial f(\varepsilon_{\mathbf{k}})}{\partial \varepsilon_{\mathbf{k}}} v_{\alpha}(\mathbf{k}) v_{\beta}(\mathbf{k}) + \frac{e^2}{\omega V} \int_0^\infty dt \, e^{i\omega t} \langle [J_\alpha(t), J_\beta(0)] \rangle^R, where the first term is the diamagnetic (paramagnetic) contribution involving the and velocities v_{\alpha} = \partial \varepsilon_{\mathbf{k}} / \partial (\hbar k_{\alpha}), f(\varepsilon) is the Fermi-Dirac distribution, and the second term is the retarded \langle \cdot \rangle^R of the total current operators J_\alpha = -e \sum_i v_{\alpha,i}. This formula separates into diamagnetic and paramagnetic parts, with the latter arising from the linear response \chi_{J_\alpha J_\beta}^R(\omega). For non-interacting electrons, it reduces to a sum over single-particle matrix elements, enabling computations in band theory. In the DC limit (\omega \to 0), the real part of the \operatorname{Re} \sigma(\omega) determines the ohmic dissipation, related to the low-frequency current fluctuations via the . For isotropic systems, the form emerges as \sigma(\omega) = \frac{ne^2 \tau}{m} \frac{1}{1 - i\omega \tau}, where \tau is a relaxation time obtained from the correlator's , illustrating how microscopic correlations yield macroscopic . This conductivity is finite in metals due to , while insulators exhibit \sigma(0) = 0. For fields, \operatorname{Im} \sigma(\omega) reflects inductive effects, with interband transitions contributing at higher frequencies in semiconductors. Applications of the Kubo extend to anisotropic materials, such as layered superconductors, where off-diagonal describe the : \sigma_{xy}(\omega) = \frac{ie^2}{\hbar \omega V} \sum_{n \neq m} \frac{f_n - f_m}{\varepsilon_n - \varepsilon_m} \langle n | v_x | m \rangle \langle m | v_y | n \rangle. Quantum Hall , quantized as \sigma_{xy} = \frac{e^2}{h} \nu for filling \nu, follows directly from this in two-dimensional gases under , highlighting topological aspects. Numerical evaluations, often via diagrammatic techniques or exact , confirm these predictions in correlated systems like the .

References

  1. [1]
    [PDF] 4. Linear Response - DAMTP
    Linear Response. Page 1. 4. Linear Response. The goal of response theory is to figure out how a system reacts to outside influences.
  2. [2]
    [PDF] Chapter 3 Linear Response Theory
    The interaction between the neutron magnetic moment and the local spin den- sity leads to a term on the Hamiltonian which is linear in S(x, t). Neutron.
  3. [3]
    [PDF] Chap 3 Linear response theory
    Eq (16) is called the Kubo formula, and χABα is called the response function. Be aware that the operators are written in the interaction picture. II. DENSITY ...
  4. [4]
    Statistical-Mechanical Theory of Irreversible Processes. I. General ...
    To add this web app to the home screen open the browser option menu and tap on Add to homescreen.
  5. [5]
    Linear response theory for open systems: Quantum master equation ...
    Feb 27, 2017 · A linear response function that characterizes how a relaxation process deviates from its intrinsic process by a weak external field is obtained ...
  6. [6]
  7. [7]
    [PDF] Correlation functions and linear response theory
    General idea (Onsager): Disturbance in a system created by a weak external perturbation decays in the same way as a spontaneous fluctuation in equilibrium. ❑ ...
  8. [8]
    [PDF] Time-Dependent Statistical Mechanics 9. Linear response theory in ...
    Oct 8, 2009 · For classical mechanics, an autocorrelation function is symmetric under change of the sign of time. An expression for the time derivative of ...
  9. [9]
    11.1: Classical Linear Response Theory - Chemistry LibreTexts
    Oct 31, 2021 · We will use linear response theory as a way of describing a real experimental observable and deal with a nonequilibrium system.Properties of the Response... · The Susceptibility · Nonlinear Response Functions
  10. [10]
    [PDF] What did Kramers and Kronig do and how did they do it?
    Mar 30, 2010 · Kramers–Kronig relations are translations into somewhat cryptic frequency language of statements clearer in time language. This translation is ...<|control11|><|separator|>
  11. [11]
  12. [12]
    [PDF] The fluctuation-dissipation theorem
    Abstract. The linear response theory has given a general proof of the fluctuation- dissipation theorem which states that the linear response of a given ...
  13. [13]
    [PDF] LINEAR RESPONSE THEORY 3.1 The generalized susceptibility
    Feb 3, 2018 · We derive the Kubo formula for the response function and, through its connection to the dynamic correlation function, which determines the ...
  14. [14]
    [PDF] The Kubo formula of the electric conductivity
    . The time evolution of the system is given by the Liouville-von Neumann equation: ∂ˆρ(t). ∂t. = 1 ih. [ ˆH(t), ˆρ(t)]. (2). We split the density operator ˆρ ...
  15. [15]
  16. [16]
  17. [17]
  18. [18]
    [PDF] Boltzmann Equation and Kubo Formula Branislav K. Nikolić
    ❑For neutral or ionized gases, the BE and its range of validity, can be directly derived from the Hamiltonian for the classical gas of molecules.
  19. [19]
    An update on the nonequilibrium linear response - IOPscience
    Jan 3, 2013 · The unique fluctuation–dissipation theorem for equilibrium stands in contrast with the wide variety of nonequilibrium linear response formulae.
  20. [20]
    Statistical-Mechanical Theory of Irreversible Processes. I.
    Statistical-Mechanical Theory of Irreversible Processes. I. General Theory and Simple Applications to Magnetic and Conduction Problems. By Ryogo KUBO.
  21. [21]
    [PDF] Chap 3 Linear response theory
    Comparing with the Kubo formula, we find the following replacement necessary,. A. → ρe,. (17). B. → ρe, f. → φext. The Kubo formula gives δhρe(x)i = Z dx0χρe ...